You are on page 1of 14

Proceedings of the 2012 9th International Pipeline Conference IPC2012 September 24-28, 2012, Calgary, Alberta, Canada

IPC2012-90459

CRACK-LIKE DEFECTS IN PIPELINES: THE RELEVANCE OF PIPELINE-SPECIFIC METHODS AND STANDARDS


Andrew Cosham Atkins Newcastle upon Tyne, UK Phil Hopkins Penspen Newcastle upon Tyne, UK Brian Leis Battelle Memorial Institute Columbus, Ohio, USA

ABSTRACT Oil and gas transmission pipelines have a good safety record; however, like any engineering structure, pipelines do occasionally fail. The main causes of pipeline failures in North America and Europe are defects, such as damage due to external interference, corrosion defects, or material defects. Consequently, the pipeline industry developed its own pipeline-specific methods for assessing the significance of these defects. These pipeline-specific methods have their origins in fracture mechanics, but the complexity of the underlying failure mechanisms was addressed through empiricism. Over the intervening years, more accurate pipeline-specific methods have been developed that better model the underlying failure mechanisms. The toughness of line pipe steels is typically characterised in terms of the Charpy V-notch impact energy. This is a qualitative measure of toughness. The pipeline-specific methods use empirical correlations between Charpy V-notch impact energy and fracture toughness. In parallel to the development of the pipeline-specific methods, fracture mechanics has been generalised in standards such as BSI 7910 : 2005 and API 579-1/ASME FFS-1, using the failure assessment diagram (FAD). The pipeline-specific methods and the methods in these general (generic) standards have common roots. The pipeline industry has used its pipeline-specific methods for 50 years, as there was little need to use the generic methods. This was because most of the defects detected in pipelines were corrosion or damage (e.g. dents), where pipeline-specific methods were well-researched and validated. The increasing sophistication of in-line inspection methods (intelligent pigs) in the pipeline industry mean that cracks in pipelines are now easily detected. Additionally, recent failures caused by crack-like defects (e.g. the San Bruno failure in the USA in 2010), mean that the industry needs guidance on how to assess these crack-like defects. There is little or no guidance on pipeline-specific methods for crack-like defects. Therefore, the

industry has adopted the generic methods, without any evaluation of their relevance or accuracy. The similarities and differences between the generic and pipeline-specific methods for crack-like defects in pipelines are illustrated in this paper through comparison with published fullscale test data for pipes containing notches and cracks. The significance of correlations between Charpy V-notch impact energy and fracture toughness is highlighted. In conclusion, a commentary is given on how and when to take advantage of either approach. The critical gaps in the existing methods are also identified. INTRODUCTION Pressure systems, such as pipelines, utilise steels and other materials to contain fluids across a broad range of temperatures and pressures. In this paper, the focus is on large diameter pipelines, as can be found in the nuclear and electric power generation, and in hydrocarbon transportation. An important part of integrity management is the assessment of the severity (significance) of defects whether dealing with defects inadvertently introduced during pipe manufacture, or fabrication and construction, or those that have developed in-service, due to degradation. Defects are typically found pre-service or in-service using one of the ever-improving inspection technologies, like in-line inspection (ILI), also known as smart or intelligent pigs. Defects that develop in-service tend to either be blunt (e.g. galvanic corrosion) or crack-like (e.g. stress corrosion cracking). Crack-like defects are planar defects, i.e. having no significant volume and a sharp tip. Crack-like defects tend to be axially-oriented in transmission pipelines, whereas the orientation of blunt defects is determined by factors like anti-corrosion coating faults (holidays), which provide the access for the environment, or gravity (e.g. bottom-of-line corrosion). The behaviour of blunt defects such as corrosion is generally both well understood, and well documented.

Copyright 2012 by ASME

The pipeline industry developed pipeline specific methods for assessing defects in pipelines. The most common types of defects in pipeline were and are corrosion defects and damage due to external interference (e.g. dents and gouges). The pipeline specific methods are based on methods for assessing crack-like defects. Extensive research and full-scale testing was conducted to validate the application of the pipeline specific methods to the types of defects typically found in pipelines. These methods have been widely used over the past 50 years. In the intervening years, fracture mechanics methods have been codified, e.g., BS 7910 : 2005 [1] and API 579-1/ASME FFS-1 2007 [2]. These are generic defect assessment methods. Standardisation has many benefits. However, standard practices developed in one industry might not be directly applicable to another industry. The pipeline industry has mainly used its pipeline specific methods, having had little need to use the generic methods. However, the generic methods have gained popularity in the industry, particularly for the assessment of crack-like defects. The increasing sophistication of in-line inspection methods in the pipeline industry mean that cracks in pipelines are now easily detected. Additionally, recent failures caused by cracklike defects (e.g. the San Bruno failure in the USA in 2010 [3]), mean that the industry needs guidance on how to assess these crack-like defects. There is little or no guidance on pipelinespecific methods for crack-like defects (although the NG-18 equations, see below, are, in fact, applicable to crack-like defects). Therefore, the industry has adopted the generic methods, without any evaluation of their relevance or accuracy. This paper first summarises the development of the pipeline-specific methods. The similarities between these specific methods and the generic methods in BS 7910 and API 579-1/ASME FFS-1 (hereafter referred to as API 579-1), are illustrated. The specific and generic methods are then compared using: 1) full-scale test data used to develop the pipeline-specific methods; and, 2) other full-scale tests on longitudinally-orientated crack-like defects in the pipe body of large-diameter pipelines. THE HISTORY OF DEFECT ASSESSMENT IN PIPELINES The first intelligent pig, a calliper tool for detecting dents, was introduced by TD Williamson in 1959 [4]. At the same time, Pan-American Petroleum was developing a tool using magnetic flux leakage (MFL) technology to detect metal loss due to corrosion. The first MFL tool, the LINALOG 90, was introduced by Tuboscope in 1965 [4]. The first generally-available report on the significance of dents in pipelines was published in 1981 [5]. The first edition of ASME B31G Manual for Determining the Remaining Strength of Corroded Pipelines [6] was published in 1984. It was based on work conducted by the Battelle Memorial Institute under the NG-18 project in the late 1960s and early 70s, e.g., Kiefner & Duffy (1973) [7]. Similarly, British Gas

first published work on the significance of metal loss in 1974 and on dents in 1981 [8-9]. Consequently, it is noteworthy that: The first in line inspection tool capable of detecting dents in a pipeline predates the development of pipeline-specific methods for assessing the results of the inspection by approximately 20 years. The first tool for detecting corrosion predates the pipelinespecific assessment methods by approximately 10 years. This race between inspection methods and assessment methods continues: the first crack-detection tool in pipelines was the elastic wave pig developed in the 1980s. However, there is, as yet, no generally-recognised pipeline-specific method for assessing crack-like defects. Fracture mechanics methods have been available since the 1950s, albeit mainly for elastic material. These methods were extended into plasticity-corrected linear-elastic fracture mechanics and then non-linear fracture mechanics, for application to elastic-plastic materials in the late 1950s to 1970s. However, the methods were not directly applicable to thin-walled, low constraint structures such as pipelines. Pipeline-specific methods needed to be developed. The NG18 Equations (see below) were developed in the 1960s and 1970s to assess longitudinally-orientated defects in pipelines, and have their roots in the assessment of crack-like defects. .Nowadays, cracks and seam weld defects in pipelines are typically assessed using the generic methods, such as BS 7910 or API 579-1. This is because there is no generally accepted pipeline-specific method for assessing crack-like defects in pipelines, and also because the generic methods are codified. The generic methods in these standards are being applied to thin-walled pipelines and thick-walled pressure vessels. It is therefore relevant to ask whether something is being lost in the application of the generic methods to pipelines. Perhaps these methods are too conservative, or inappropriate? These questions can partly be answered by comparing the pipeline-specific methods with BS 7910 and API 579-1. BS 7910 : 2005 BS 7910, Guide to methods for assessing the acceptability of flaws in metallic structures, outlines methods for assessing the acceptability of flaws in all types of structures and components. BS 7910 has its origins in PD 6493 : 1980 [10,11]. Acceptance is based on the concept of fitness for purpose. By this principle, a particular fabrication is considered to be adequate for its purpose, provided the conditions to cause failure are not reached. An engineering critical assessment is used to decide whether or not a flaw is acceptable on the basis of fitness for purpose. Planar flaws, non-planar flaws and shape imperfections may be assessed using the document. It is, however, primarily concerned with the assessment of planar flaws, i.e. cracks. Therefore, it is commonly used when crack-like defects are detected in pipelines.

Copyright 2012 by ASME

The assessment of crack-like flaws is based on a twoparameter failure assessment diagram. The failure assessment diagram considers elastic-plastic fracture and plastic collapse. API 579-1/ASME FFS-1, 2007 API 579-1/ASME FFS-1, Fitness-For-Service, provides guidance for assessing the fitness-for-service of in-service flaws or damage in pressurised equipment. The first edition, published as API 579 in 2000, was developed for application in the refinery and petrochemical industry, but it was also widelyused in other industries. The second edition, API 579-1, was published as a joint API and ASME standard in 2007, so as to be applicable to all pressured equipment designed and constructed to an API or ASME standard. The assessment of crack-like features in API 579-1 is similar to BS 7910. It uses a failure assessment diagram. PIPELINES AND FRACTURE MECHANICS Fracture mechanics developed largely in a linear-elastic framework (e.g. Griffith (1920) and Irwin (1957) [12,13]). It was difficult to apply this linear-elastic fracture mechanics to the thin sectioned, low strength, ductile steels in use in the pipeline industry. The pipeline industry began research on the significance of defects in pipelines in the early 1960s. This led to the log-secant analysis of longitudinally-orientated cracklike defects in cylinders subject to internal pressure, known as the NG-18 Equations [14-25]. The methods are based on experiments and theory. Fullscale and model-scale experiments helped the researchers at that time to overcome a number of limitations in (the then) contemporary fracture mechanics theory: 1. Linear-elastic fracture mechanics theory was not able to accommodate the ductility of line pipe steels. The yielding associated with the failure of a defect in line pipe could not be included in models in the 1950s and 1960s. Relatively thin-walled pipelines have low constraint. Fracture toughness varies with constraint: the lower the constraint, the higher the toughness. The standard fracture mechanics specimens used to measure toughness have a high constraint. Fracture mechanics theory considers sharp defects, i.e. cracks. Many of the defects in pipelines are blunt (e.g. corrosion). Blunt defects do not fail at the low pressures that cracks do. Sharp defects blunt in ductile steels. Toughness of line pipe has traditionally been measured using the Charpy V-notch (CVN) impact test specimen. The impact energy is a qualitative measure of toughness. It is not representative of the toughness of the structure unless it is validated and calibrated in full-scale tests. Quantitative measures of fracture toughness, such as the crack tip opening displacement or J-integral, have failed to gain popularity in the pipeline industry (except in welding), as their use would still need to be validated using full scale test data.

There remain limitations on the application of contemporary fracture mechanics to pipelines. Toughness is still measured using the Charpy V-notch (CVN) impact test. The assessment of crack-like defects in the pipe body using the generic methods is therefore immediately limited by the absence of a quantitative measure of fracture toughness. Additionally, constraint-based fracture mechanics is not yet well codified in the standards. THE NG-18 EQUATIONS The Battelle Memorial Institute developed the semiempirical NG-18 Equations in the 1960s and 70s. The equations are failure criteria for longitudinally orientated, through-wall and part-wall defects in a cylinder subject to internal pressure. The final forms of the failure criteria are described by Maxey et al. (1972) [23]. Fracture-controlled and flow-stress controlled forms were developed1, based on the two types of failure that were observed in the full-scale tests: 1. Fracture-controlled failures, for which a measure of the toughness was required to predict the failure stress; and, 2. Flow-controlled failures for which a measure of the toughness was not required to predict the failure stress. The NG-18 equation for a part-wall defect is given in equation (1a), or re-arranged in terms of the hoop stress at failure, in equation (1b)2. Equation (1) is the toughness dependent form. If the toughness is high, it reduces to the flow stress dependent form, given in equation (2). These failure criteria are semi-empirical. The flow stress, and the relationship between the fracture toughness (stress intensity, K or strain energy, G) and the upper shelf Charpy Vnotch impact energy were defined using the results of 92 full scale tests on vessels containing through-wall defects [23,24]. Hahn et al. (1969) [20] plotted the Charpy V-notch impact energy of 16 tests on through-wall defects against the value of G determined from the results of the tests (i.e. the tests were used as large fracture toughness test specimens), repeated in Fig. 4. The 16 tests were all toughness dependent failures3. The impact energy of the 16 data points ranges from 13.666.4 J (10-49 ft.lbf). A linear relationship between the energy density and G was assumed, see Fig. 4 and equation (1). Note that the fracture toughness is a post-yield form of the planestress linear elastic fracture mechanics crack driving force, K, expressed as an energy release rate, G, to give consistent units (it could equally be expressed as J). The scatter in the data indicates that the linear relationship is an assumption, but it met the needs of the researchers at the time. The correlation that was at the time considered linear, in fact becomes quite nonlinear as the toughness of the steel increases [26,27]. The NG-18 equations are representative of the specific methods for assessing longitudinally-orientated, crack-like
Also referred to as toughness dependent and flow stress dependent, respectively. 2 Equations 1a and 1b are given for units of millimeters, Newtons and Joules. 3 The failure stress was less than 0.8 .
1

2.

3.

4.

Copyright 2012 by ASME

defects. The NG-18 equations are applicable to crack-like defects. There are other, more recent, methods. The Ductile Flaw Growth Model (DFGM), and the related Pipe Axial Flaw Failure Criterion (PAFFC), were developed by the Battelle Memorial Institute in the mid 1990s [28,29]. These are an update to the NG-18 equations. CorLAS (Corrosion Life Assessment Software) was developed by CC Technologies for assessing stress corrosion cracking in pipelines [30-32]. In approach, it has some similarities with the DFGM, but also includes a method for assessing the actual profile of a cracklike defect in terms of an effective semi-elliptical crack. The Link to the Generic Methods The NG-18 equations are failure criteria for through-wall and part-wall defects in a cylinder subject to internal pressure. They can be used to assess the significant of crack-like defects. Similarly, BS 7910 and API 579-1 can be used to assess cracklike through-wall and part-wall defects. The methods in BS 7910 and API 579-1 use a failure assessment diagram. The two approaches (the specific and generic methods) are different, but the differences are only superficial. The NG-18 equations can also be expressed in terms of a failure assessment diagram, as illustrated below. (1a) (1b) (2)

r = 0.003-4 in.

r = 0.005 in.

Fig. 1 Through-wall and part-wall defects

(3a) (3b) (3c) (3d) (3e) (3f)

In equations (3a-3h), the through-wall equation is rearranged into the form of a failure assessment diagram. The resulting failure assessment diagram is mathematically identical to that of the Level 2 method in PD 6949 : 1991, i.e. a logsecant formulation. The NG-18 equations and the Level 2 failure assessment diagram are both based on the strip yield model [33-35]. It is therefore unsurprising that the throughwall form of the NG-18 equations can be directly related to this failure assessment diagram. Similarly, the part-wall form of the NG-18 equations can also be related to this form of the failure assessment diagram. A direct relationship can then be drawn between the NG-18 equations and the failure assessment diagrams in BS 7910 and API 579. The only differences are then in the shape of the FAD and the equations that are used to calculate (using the terminology of BS 7910) the applied stress intensity factor and the reference stress. Indeed the flow stress dependent forms of the NG-18 equations have been incorporated into BS 7910 and API 579. Having established that the NG-18 equations and the methods in API 579 and BSI 7910 are essentially the same, it is then instructive to compare predictions made using these methods to the full-scale test data, i.e. to compare the specific and generic methods. The comparison (see below) shows that there are significant differences in the predictions. The generic methods are conservative compared to the NG-18 equations. The main reason for this is their differing correlations between fracture toughness and Charpy V-notch impact energy. THE NG-18 THROUGH-WALL AND PART-WALL TEST DATA The NG-18 equations were developed using the results of 92 tests of through-wall defects and 48 tests of part-wall defects. The NG-18 equations and BS 7910 are compared using these same tests. Part 9 of API 579-1 is similar to BS 7910, so BS 7910 is representative of the generic methods. The tips of the through-wall notch were made with a jewellers saw cut. This produced a defect tip with a radius of 0.08 to 0.10 mm (0.003 to 0.004 inch) [17]. The part-wall defects were milled V-shaped notches. The radius at the tip of the V-shaped notch was approximately 0.127 mm (0.005 inch) [17,24]. The through-wall and part-wall defects are crack-like defects. The radius at the tip of the defects is at least an order of magnitude smaller than the criterion given in API 579-1 for

(3g)

(3h)

Copyright 2012 by ASME

predicted failure stress (sq/sy), percent

120
TW NG-18 (toughness dependent) BS 7910 : 2005 API 579-1

100

80
60 40 20 0 0 20 40 60 80 100 120 actual failure stress (sq/sy), percent

Fig. 2 Comparison of predicted and observed failure stress for through-wall defects

predicted failure stress (sq/sy), percent

PW NG-18 (toughness dependent) BS 7910 : 2005

predicted failure stress (sq/sy), percent

120
100

120
PW NG-18 (flow stress dependent) BS 7910 : 2005
CVN less than 21 J (16 ft.lbf) CVN not reported

100

80
60 40
U-shaped notch

80
60 40 20 0 0

U-shaped notch

20 0 0
2c

2c

20 40 60 80 100 120 actual failure stress (sq/sy), percent

20 40 60 80 100 120 actual failure stress (sq/sy), percent

a) toughness-dependent NG-18 equation

b) flow stress dependent NG-18 equation

Fig. 3 Comparison of predicted and observed failure stress for part-wall defects

Copyright 2012 by ASME

NG-18

BS 7910 API 579-1

A correlation is given for steels on the upper shelf. It is a curve-fit between the upper shelf Charpy V-notch impact energy and the fracture toughness inferred from full-scale burst tests on pipes containing longitudinal through-wall defects (see Fig. 4), i.e. the full-scale test is the fracture toughness test specimen. A correlation is given for steels exhibiting lower shelf and transitional behaviour (clause J.2.1) and a correlation is given for steels on the upper shelf (clause J.2.4). Two correlations are given for steels exhibiting transitional behaviour (and, by inference, steels on the lower shelf) (clause F.4.5.2b) and two correlations are given for steels on the upper shelf (clause F.4.5.2a).

Note: 1. The correlations summarised above do not require a Charpy V-notch transition curve. BS 7910 : 2005 and API 579-1 also give correlations based on the master curve approach. These correlations do require a transition curve. 2. The correlations in BS 7910 : 2005 and API 579-1 are lower bound correlations. They are based on curve-fits between the Charpy V-notch impact energy and a fracture toughness determined using a standard fracture toughness test specimen. Table 1 Correlations between Charpy V-notch impact energy and fracture toughness

assessing a groove as a crack-like defect4. The geometry of the through-wall and part-wall defect is illustrated in Fig. 1. Cracks or crack-like defects in moderate to high toughness materials will blunt before failure, so there is no significant difference between machined defects, with a small root radius, and fatigue pre-cracked defects, when such defects are pressurised to failure. The actual failure stress is plotted against the failure stress predicted using the NG-18 equations, BS 7910 and API579-1 in Fig. 2, for through-wall defects, and predicted using the NG-18 equations and BS 7910 in Fig. 3, for part-wall defects5. The reported material properties (yield and tensile strength, and upper shelf Charpy V-notch impact energy) are used. The Charpy V-notch impact energy (2/3 specimens size) for the through-wall tests is in the range 13.6-90.9 J (10-67 ft.lbf), and for the part-through-wall tests it is 13.6-46.1 J (10-34 ft.lbf). Through-Wall Defects Fig. 2 shows very good agreement between the experimental data and the NG-18 equations. This is to be expected; the NG-18 equations were developed using this data (but note that the equations are not a curve-fit to the data). The predictions made using BS 7910 and API 579-1 are relatively conservative. The data points do, however, follow a consistent trend. The two data points off the trend, with an actual failure stress of approximately 120% of yield are both tests of 25.4 mm (1 inch) long defects. These are the shortest defects in the data set. The predictions made using API 579-1
API 579-1 uses the root radius at the root of a flaw to distinguish between groove-like flaws and crack like flaws 5 In the predictions made using the NG-18 equations, the flow stress is equal to the yield plus 10 ksi (approximately 69 N.mm-2) and the three term Folias factor is used. The actual length of the part-wall defects is used rather than the (shorter) equivalent length. In the predictions made using BS 7910, the factor of 1.2 on M in Clauses P.4.2.1 and P.4.3.1 is not included. The effect of including the factor would be to increase the conservatism of the predictions. Otherwise, a Level 2A assessment is conducted in accordance with the standard. In the predictions made using API 579-1, the stress intensity factor solution for the inside surface of a through-wall crack subject to internal pressure (Gp in Table C.6) is used. Equation F.60 (Rollfe-Novak-Barsom) is used to calculate the fracture toughness. A Level 2 assessment is conducted.
4

are, relatively, less conservative than those made using BS 7910. The difference can largely be attributed to the correlations between Charpy V-notch impact energy and fracture toughness. That in API 579-1 gives a higher toughness than that in BS 7910, as illustrated in Fig. 5. Part-Wall Defects Fig. 3a uses the toughness-dependent NG-18 equations, and Fig. 3b uses the flow stress dependent equations. The scatter in the data points is greater than for the through-wall test data. This is to be expected, because the part-wall NG-18 equations are more semi-empirical than the through-wall equations. The flow stress dependent equation gives good predictions for all of the tests where the reported upper shelf Charpy V-notch impact energy is greater than 21 J (2/3 specimen size). The effect of acuity is illustrated by the results for the one test on a U-shaped notch (this is not a crack-like defect). The test is accurately predicted using the flow stress dependent equation even though the toughness is less than 21 J. This observation is further supported by the findings of more recent studies of the effect of toughness on the burst strength of corrosion defects [36]. The predictions made using BS 7910 are, again, conservative compared to NG-18. The data points do, however, follow a consistent trend. The failure assessment diagram is intended to describe failure avoidance rather than failure prediction. However, this distinction is not sufficient to explain the trends in Fig. 2, and Fig. 3. Applying BS 7910 to a through-wall defect in a cylinder subject to internal pressure results in essentially the same equations as the NG- 18 equations (and similarly for partwall defects, with the exception of the equation for the stress intensity factor solution). The Level 2A FAD in BS 7910 is slightly different to what was the Level 2 FAD in PD 6949 : 1991, but again this is not sufficient to explain the trends. Therefore, given that the pipeline-specific and generic methods are essentially the same, why are the generic methods overly conservative when used to predict the failure of crack like-defects in pipelines?

Copyright 2012 by ASME

THE RELATIONSHIP BETWEEN FRACTURE TOUGHNESS AND CHARPY V-NOTCH IMPACT ENERGY The NG-18 equations incorporate a linear relationship between fracture toughness (expressed in terms of G) and the

10
strain energy release rate (G), ksi.in 0.5

8
9% Ni alloy

9% Ni alloy

0 0 2 4 6 12CVN/A, ksi.in0.5 8 10

Fig. 4 Relationship between the strain energy release rate and the upper shelf Charpy V-notch impact energy, after Hahn et al. (1969)

upper shelf Charpy V-notch impact energy density, see Fig. 4. The relationship can also be expressed in terms of the stress intensity factor, K, and the Charpy V-notch impact energy, as in Fig. 5. The impact energy is the upper shelf energy (i.e. 100 percent shear area). The Charpy V-notch impact energy is a qualitative measure of toughness. The methods in BS 7910 and API 579-1 for assessing crack-like flaws require a quantitative measure of fracture toughness (e.g. the crack tip opening displacement or the J-integral). Correlations between fracture toughness and Charpy V-notch impact energy have been developed. These correlations tend to be conservative. Annex J of BS 7910 and Annex F of API 579-1 give correlations between impact energy and fracture toughness for use when fracture toughness data is not available. The available correlations are summarised in Table 1. The relationships for upper shelf steels are plotted in Fig. 56. It can be readily seen that these correlations are very conservative in comparison to the relationship used in the NG-18 equations. The correlations in BS 7910 and API 579-1 are based on fracture toughness tests using standard fracture toughness test specimens. The correlation in the NG-18 equation is based on tests on through-wall defects, as summarised in Fig. 4. It is therefore to be expected that the correlations for upper shelf steels in BS 7910 and API 579-1 are very conservative in comparison. The correlation embedded in the NG-18 equations is only applicable to upper shelf steels. Older steels may not exhibit upper shelf behaviour. For steels that exhibit transitional or lower shelf behaviour, the correlations given in BS 7910 and API 579-1 must be used. The generic methods are relatively conservative when compared to the specific methods, as illustrated in Fig. 2 and Fig. 3, because of the correlations between upper shelf Charpy V-notch impact energy and fracture toughness are very conservative when applied to line pipe steels. Line pipe steels are typically ductile and tough, and pipelines are thin-walled in comparison to some other pressurised structures. The conservatism in the correlations for upper shelf steels has two, related implications for the application of the generic methods. Firstly, the predicted failure stress of toughness dependent failure is lower. Secondly, some failures are predicted to be toughness dependent, whereas they are in fact flow stress dependent. The semi-empirical NG-18 equations were developed using the 92 tests of through-wall defects and the 48 tests of part-wall defects presented in Fig. 2 and Fig. 3. A comparison between the specific and generic method using only this test data is somewhat limited. The NG-18 equations would be

500
Hahn et al. (1969)

fracture toughness, MPa.m0.5

400
9% Ni alloy

300

200

Annex F, ASME FFS-1 2007


(Barsom & Rollfe, 1999)

100
Annex J, BS 7910 : 2005

0 0 20 40 60 80 full-size CVN impact energy, J 100

Fig. 5 Relationship between upper shelf Charpy V-notch impact energy and fracture toughness in the NG-18 equations, and in Annex J of BS 7910 : 2005 and Annex F of API 579-1

Equation (J.1) from BS 7910 is plotted for a wall thickness of 6.35 and 12.7 mm. Equation (J.5) is independent of the wall thickness. Equation (F.60) from API 579-1 is plotted for yield strengths of 245 and 450 N.mm-2 (i.e. approximately Grade B to X65).

Copyright 2012 by ASME

predicted failure stress (sq/sy), percent

120
NG-18 (TW or PW) BS 7910 : 2005
Garwood et al. (1981) 14 J (10 f t.lbf ) Leis (2001) Hosseini et al. (2010)
TW

TW

100

80
60 40 20 0 0

20 40 60 80 100 120 actual failure stress (sq/sy), percent

Fig. 6 also includes tests for part-wall defects reported by Leis (2001) [38] and Hosseini et al. (2010) [39]. Leis (2001) describes the results of tests of four vessels containing part-wall defects, all machined V-shaped notches (the root radius was approximately 0.05 mm). The full-size upper shelf Charpy Vnotch impact energy the test vessels was 20, 141, 157 and 292 J (15, 104, 116 and 215 ft.lbf). The impact energy of the three higher toughness line pipe is typical of modern line pipe steel, and higher than that of the data on which the correlations between Charpy V-notch impact energy and fracture toughness are based. The trends for these five tests are consistent with Fig. 3. The predictions using BS 7910 are better for the tests in the higher toughness steels, because the estimated fracture toughness is higher. Hosseini et al. (2010) describe four tests of fatigue pre-cracked defects. The shape of the defects is approximately semi-elliptical. The predictions using the NG-18 equations are more conservative than in some other tests because of the profile of the defects7. These tests highlight the importance of another issue, the profile of the defect. Cracks and crack-like defects are not typically flat-bottomed. COMPARISON USING FIELD DATA Rothwell and Coote (2009) [40] compare predictions made using the toughness-dependent NG-18 equations, PAFFC, CorLAS and BS 7910 against field and hydrostatic test failures: eighteen caused by cracks due to stress corrosion cracking (SCC), one due to a hook crack, and one due to a fatigue crack8. The actual yield and tensile strength, and Charpy V-notch impact energy were used. The data includes line pipe steels on the upper shelf and in the transitional region. Information on the flaw shape was available for eight of the failures. Considering firstly only the predictions made using the NG-18 equations and BS 7910 (the two other specific methods are subsequently revisited below), both sets of predictions are conservative. The NG-18 predictions are marginally less conservative. The difference between the NG-18 and BS 7910 predictions reported in Rothwell and Coote (2009) are less pronounced than observed in Fig. 2, Fig. 3 and Fig. 6 because the correlation between Charpy V-notch impact energy and fracture toughness embedded in the NG-18 equations was used in the BS 7910 predictions. This illustrates the issues around defining a common basis on which to compare different assessment methods. A further complication is that, although
7 The predictions using the NG-18 equations would be more accurate if the profile of these semi-elliptical flaws was taken into account. Hosseini et al. (2010) report that predictions of these tests made using API 579-1 are more accurate than those made using the NG-18 equations. The reference stress solution in API 579-1 for a semi-elliptical (parabolic) flaw includes a factor of 0.67 to account for the profile of the flaw. Including this factor in the NG-18 equations (as in ASME B31G) gives predictions that are lower than the actual failure stress, but higher than those of the Level 3 analysis reported in Hosseini et al. (2010). 8 Rothwell and Coote (2009) also compared predictions using the same 47 tests of part-wall crack-like defects as are in Fig. 3. However, in their predictions using BS 7910, Rothwell and Coote used the correlation between Charpy Vnotch impact energy and fracture toughness embedded in the NG-18 equations, not that in Annex J. If that correlation is used then the predictions made using BS 7910 are significantly improved compared to those presented in Fig. 3.

Fig. 6 Comparison of predicted and observed failure stress for some other though-wall and part-wall defects

expected to give reasonably accurate predictions of these tests. It is therefore instructive to also consider other test data and field data. COMPARISON USING OTHER TEST DATA A number of other tests on through-wall and part-wall defects in line pipe steels have been published since the early 1970s. The significance of the correlation between Charpy Vnotch impact energy and fracture toughness indicates that the most valuable test data is that for which the fracture toughness of the line pipe steel has been measured. However, little such data has been published for line pipe steels. Tests on thick-wall pressure vessels steels are not relevant to typical line pipe steels. Garwood et al. (1981) [37] report two burst tests in X56 line pipe steel, one through-wall defect and one part-wall defect, for which the J-integral of the line pipe steel was measured. A more detailed review of the literature will likely identify more relevant tests with both Charpy V-notch and fracture toughness data than considered here, but the Garwood et al. (1981) data is sufficient to illustrate the trends. The actual failure stress is plotted against the failure stress predicted using the NG-18 equations and BS 7910 in Fig. 6, for the two tests reported in Garwood et al. (1981). The (upper shelf) fracture toughness (Jm) measured in the TS orientation is used for the part-wall defect and that measured in the TL orientation is used for the through-wall flaw. The toughness was measured using single edge notch bend specimens (i.e. high constraint). The predictions made using BS 7910 are closer to the actual failure stress because the fracture toughness is higher than that implied by the correlation between toughness and Charpy V-notch impact energy. The agreement would be further improved by using lower constraint test specimens.

Copyright 2012 by ASME

some of the line pipe steels were not on the upper shelf, the predictions are conservative. The correlation embedded in the NG-18 equations (and used in both the BS 7910 and CorLAS predictions) is notionally only applicable to upper shelf steels, but has been applied to steels that exhibit transitional behaviour. That the predictions are conservative is due to other conservatisms in the assessment methods.

5
Gp (Table C-6, API 579-1) M 4
t/Ri = 0.01667

t/Ri = 0.05

t/Ri = 0.1

1 0 100

762 mm (30 in.)

200 300 crack length, 2a

400

500

Fig. 7 Comparison of the stress intensity factor solution for a longitudinal, through-wall crack in a cylinder in Annex C of API 579-1 and the Folias factor, M

5
Mt (Equation (D.8), API 579-1) M 4
t/Ri = 0.01667 t/Ri = 0.05

t/Ri = 0.1

1 0 100

762 mm (30 in.)

200 300 crack length, 2a

400

500

STRESS INTENSITY FACTOR AND REFERENCE STRESS SOLUTIONS BS 7910 and ASME FFS-1 give stress intensity factor solutions and reference stress solutions for various geometries. The different solutions are not the cause of the large differences between the specific and generic methods observed in Fig. 2, Fig. 3 and Fig. 6, as is shown below. The case of a longitudinally orientated through-wall defect in a cylinder subject to internal pressure is one of the simplest geometries. It is therefore instructive to consider this case. The through-wall NG-18 equations use the Folias factor, M, as both (in the terminology of BS 7910 and API 579-1) the stress intensity factor solution and the reference stress solution. For curved shells under internal pressure, BS 7910 also uses the Folias factor, albeit a more conservative two-term expression rather than the three-term expression used in the NG-18 equations. Fig. 7 and Fig. 8 compare the Folias factor with the respective solutions given in Annexes C and D of API 579-1. API 579-1 is a more recent document than BS 7910, so the solutions might reasonably be expected to be better. The stress intensity factor solution, for internal pressure only, in Fig. 7 is plotted for t/Ri=0.01667-0.1, equivalent to D/t=122-22. A diameter of 762 mm (30 in.) is assumed because over half of the original 92 full-scale tests were conducted on pipe of this diameter. The length of most of the through-wall defects in the tests was less than 350 mm. The stress intensity factor calculated using Annex C of API 579-1 is, in most cases, slightly smaller than the Folias factor9. The stress intensity factor calculated using the solution in Annex M of BS 7910 is slightly larger than that in API 579-1. The reference stress solution in Fig. 8 is also plotted for t/Ri=0.01667-0.1 and a diameter of 762 mm (30 in.), for consistency with Fig. 7. The reference stress solution in API 579-1 is very similar to the Folias factor. The reference stress solution in Annex P if BS 7910 is the Folias factor, apart from a multiplying factor of 1.2 (that has not been included in Fig. 2, Fig. 3 or Fig. 6). The trends in Fig. 7 and Fig. 8 demonstrate that the overconservatism evident in Fig. 2 is not due to the stress intensity factor solutions and reference stress solutions. API 579-1 will give less over-conservative predictions than BS 7910, because the stress intensity factor will be slightly lower. If the Hahn et al. (1969) correlation between Charpy V-notch impact energy and fracture toughness is used, then the predictions using BS 7910 or API 579-1 will be less over-conservative (as illustrated in Rothwell and Coote (2009)). In the limit, as the effective fracture toughness increases (perhaps through the use of constraint-based fracture mechanics), the predictions using the NG-18 equations, and BS 7910 or API 579-1 will be similar because the reference stress (collapse) solutions are essentially the same. The trends can be partially generalised to the case of a longitudinally orientated part-wall defect. The reference stress
9

Fig. 8 Comparison of the reference stress solution for a longitudinal, through-wall crack in a cylinder in Annex D of API 579-1 and the Folias factor, M

Mt, M

G p, M

Annex C of API 579-1 gives solutions for the stress intensity at the inside surface and the outside surface. That for the inside surface is plotted in Fig. 7 and used in the predictions summarised in Fig. 2. It gives (relatively) less conservative predictions.

Copyright 2012 by ASME

solutions in Annex D of API 579-1 (the local limit load solution) and Annex P of BS 7910 are very similar to the flow stress dependent form of the part-wall NG-18 equations. The flow stress dependent form reasonably describes the test data, at least where the 2/3 Charpy V-notch impact energy (upper shelf) exceeds 21 J, see Fig. 3 and Fig. 6. In the limit, as the effective fracture toughness increases, the predications of all of the methods will be similar. Through-wall and part-wall defects in modern line pipe steel fail are flow stress dependent (flow controlled). The flow stress dependent form of the part-wall NG-18 equations can be used if the impact energy is high enough. The generic methods, i.e. BS 7910 and API 579-1 are conservative compared to the NG-18 equations because they predict a toughness-dependent failure rather than a flow stress dependent failure (i.e. failure is predicted when Lr is less than Lr max). Less conservative correlations between Charpy V-notch impact energy and fracture toughness, such as that due to Hahn et al. (1969), reduce this over-conservatism. DFGM, PAFFC AND CorLAS The Ductile Flaw Growth Model (DFGM) and the related Pipe Axial Flaw Failure Criterion (PAFFC) [28,29] and CorLAS (Corrosion Life Assessment Software) [30-32] are more recently developed pipeline-specific methods. A blind round-robin comparison of the NG-18 equations, PAFFC and CorLAS using fourteen failures due to stress corrosion cracking, as part of the National Energy Boards Public Enquiry into stress corrosion cracking in Canadian pipelines [41], showed that the NG-18 equations (the logsecant method) could be very conservative, whilst PAFFC and CorLAS both gave reasonably accurate, if slightly conservative predictions. CorLAS was comparatively more accurate [41]. Rothwell and Coote (2009) [40] used the same fourteen failures together with eight additional failures, and also made predictions using BS 7910. CorLAS was again identified as giving the slightly better predictions, both when using an assumed semi-elliptical profile and the actual profile (where available) of the crack. It follows then, from the comparisons presented in Fig. 2, Fig. 3 and Fig. 6, and the comparison in Rothwell and Coote (2009) that these more recent specific methods, that are more accurate than the NG-18 equations, will be more accurate that the generic methods. The issue then is why are these more recent specific methods not used, or why have the improvements not been incorporated into the generic methods. GENERIC OR SPECIFIC METHODS? The intent here is not to say that one method is better than another method for assessing crack-like defects in pipelines. Rather, it is to discuss the issues and present a way forward. There are no fundamental differences between the specific and generic methods. Fig. 2, Fig. 3 and Fig. 6, and Rothwell and Coote (2009) [40], illustrate why the pipeline
TM

industry developed its own specific assessment methods in the past. They are more accurate than the generic methods. The accuracy of the generic methods is constrained by the conservative correlations between Charpy V-notch impact energy and fracture toughness (Annex F of API 579-1 and Annex J of BS 7910). However, measuring the fracture toughness of line pipe steel will not address the underlying issues. The standard fracture mechanics specimens used to measure fracture toughness create high constraint around the crack tip. Defects in relatively thin-walled pipelines have low constraint. Constraint-based fracture mechanics is not well codified. It is of limited practical use in pipelines because of the absence of relevant fracture toughness measurements for inservice pipelines. Cracks (sharp defects) show more of an influence of toughness, but the failure of cracks in line pipe is typically ductile. Flow stress dependent methods can be used to predict the failure of crack-like defects if the line pipe steel has a high enough toughness, as illustrated in Fig. 3b. The issue then is defining what is a high enough toughness. The NG-18 equations are simple and easy to use, and transparent. The more recent pipeline-specific methods are more complicated, and are typically software-based. They are less transparent. This lends them the air of a black-box. API 579-1 and BS 7910 are not simple, but they are more transparent. However, when implemented in software they also tend towards the black box. The generic methods are codified and may therefore be perceived to have a greater acceptance by the regulators. This may be particularly the case when assessing crack-like defects, where there is an expectation to see an engineering critical assessment conducted in accordance with a standard such as BS 7910 or API 579-1. However, as illustrated here, a more complicated method does not necessarily mean a more accurate result. The Need For Guidance The issue for the present is then not whether the specific methods should be used in preference to the generic methods, or vice versa, but rather how to move things forward. Most users of a generic method such as BS 7910 or ASME FFS-1 will use the method as it is written. Therefore, with the potential for the increasing adoption of generic methods in the pipeline industry, it is incumbent on the industry not to lose the benefits of the specific methods. This might take the form of providing guidance on how the generic methods might best be applied to pipelines, or codifying the specific methods. Examples, in a different, but related context, would be ASME B31G-2009 [42] and the EPRG guidelines on the assessment of defects in transmission pipeline girth welds [43,44]. The EPRG guidelines on the assessment of defects in transmission pipeline girth welds provide an interesting parallel. The guidelines present simple acceptance limits for defects in girth welds based on the toughness of the weld. It is well recognized that a simple application of BS 7910 (or API

10

Copyright 2012 by ASME

579-1) results in limits that are lower than those in the EPRG guidelines [45]. The limits in the guidelines were validated by extensive wide plate and full-scale testing. A more detailed assessment (Level 3) will bridge the gap between the two approaches. Similarly with the assessment of through-wall and part-wall crack-like defects. There is then a need to define the minimum toughness, measured using existing methods (i.e. Charpy V-notch impact energy), that is high enough to use the flow stress dependent methods. The principle is illustrated in Fig. 3a (toughness dependent) and Fig. 3b (flow stress dependent). The issue is then applying this to applications that involve heavier wall, higher toughness, and higher strength line pipe steels. The fracture toughness of the line pipe steel is not available for most existing pipelines, and it is of questionable value for modern, high toughness line pipe steels. The failure of defects in such steels will be flow-controlled. The issue is then not constrained-based fracture mechanics, but limit load analysis (constrained-based fracture mechanics simply serves to move the assessment point around the failure assessment diagram). The traditional approach to validating the pipelinespecific methods used to assess defects in pipelines has been full scale testing. This is expensive. Thus, a major problem with the existing pipeline-specific methods that were developed to deal with older line pipe steels is that in some cases they have not been proven for current and emerging needs. The Need for Research There is a clear need to improve the existing suite of models for defect assessment. The main problem in recent years has been a lack of interest and investment in research into pipeline defect assessment. This under-investment is in strong contrast with the financing of new intelligent pigs and their use in the field to find the defects that often must be assessed using such criteria. Consequently, the capabilities of the inspection tools are running ahead of the capabilities of the assessment methods. The wider use of generic methods does not address these underlying issues. IS IT A CRACK? The discussion so far has been concerned with predicting the burst strength of a defect of a known type and size. Reality is not as simple as this. Cracks or crack-like defects in a pipeline may first be reported following an in-line inspection. An inspection tool is, like any inspection, subject to errors. The feature found in the field might not be the same type as the feature reported by the tool. This error is characterised by the probability of identification. The size of the feature measured in the field might be larger or smaller than that reported by the tool. Metal-loss tools are, in general terms, more accurate than crack detection tools. It is easier for the intelligent pig to detect, identify and size corrosion than it is a crack. Bates el al. (2010) [46] reviewed the results of the inspection of two pipelines using crack detection tools. In the first case,

inspected in 2007, the probability of identification of crack-like features was at least 0.6. Notch-like features and weld anomalies were all correctly identified. The sizing accuracy was lower than the specified accuracy. In the second case, inspected in 2004 and 2007, the probability of identification of crack-like features was at least 0.69, but only 0.44 for notchlike features. No crack-like features were incorrectly identified as notch-like features or weld anomalies. The sizing accuracy was lower than the specified accuracy. A crack field reported in the 2007 inspection subsequently failed in 2008. The issues identified in the Bates el al. (2010) review of the inspection of two pipelines lead to the question Where are the biggest errors in the assessment? Inspection reliability and errors may well swamp any errors associated with the assessment method. The use of more accurate assessment methods does not address the issue of inspection errors. ASSESSING CRACKS IN PIPELINES The continuous improvement of crack detection in-line inspection tools means that the pipeline industry will detect more and more crack-like defects. Therefore, there is a need to have clear and safe guidelines for their assessment. There are merits in using either the specific or the generic methods. API 579 and BS 7910 are recognised standards. The specific methods are more accurate for upper shelf behaviour. The generic methods are conservative, but with the following important qualification: The generic methods are conservative (they will give small values) when used to predict the size of defect that is acceptable at the operating pressure of the pipeline. However, when used to predict the remaining fatigue life of a crack-like defect, following a pre-commissioning or in-service hydrostatic test, these methods will give non-conservative predictions of the remaining fatigue life. This is because: the predicted initial defect size (the size of defect that fails at the hydrotest pressure), will be small; whereas, the predicted final defect size (the size of defect that fails at the operating pressure) will be small and conservative. Calculations of the remaining life of crack-like defects subject to fatigue crack growth are more sensitive to the initial defect size than the final defect size. Therefore, the predicted initial defect size is, in fact, too small (the defect would not fail at the hydrotest pressure) and it is non-conservative. Predicting a larger initial defect size will result in a shorter predicted fatigue life; consequently, the generic methods will give non-conservative predictions of the remaining fatigue life. In most crack assessments, the line pipe steels toughness will either not be known, or there will be limited data. This may lead to uncertainty in determining if the line pipe will exhibit upper shelf behaviour, particularly if the crack is in a weld. In the short term, the non-expert should use API 579 or BS 7910. In the longer term, the pipeline industry needs to develop better guidance for assessing crack-like defects in pipelines, which takes advantage of the benefits of the specific methods.

11

Copyright 2012 by ASME

However, the question What method should be used to assess cracks in pipelines? is not the first question to ask. It is... What is the cause of the crack? Experts in fracture mechanics will treat cracks in pipelines with caution. This is not because the assessment is complicated, but because a crack in a structure is usually a sign that something has gone wrong, or something is wrong, or something is going wrong. It is essential to understand the cause, and consequences of any cracking detected in a pipeline before any assessment. The cracking may be an isolated incident, i.e. a one-off, but it is more likely to be a symptom of a more serious, generic problem in the pipeline. Therefore, cracks detected in a pipeline, particularly those that may grow, should be treated with extreme caution. The approach should be to prevent or detect and repair, rather than to assess. The subsequent assessment can look very simple simply run a software package that implements BS 7910 or API 579-1, and an answer is obtained. However, if the data entered into the software package is wrong then the answer will be wrong. The answer is a number, not a solution to the problem. Therefore, whatever method is used, a number of questions need to be asked before and when assessing cracks in pipelines: Is it a crack (i.e. has the inspection called it correctly?)? Is the crack really there, and the correct size (has the inspection measured it correctly?)? What is the cause of the crack (corrosion, fatigue, manufacturing, etc.)? Who is competent to assess the crack? Is it worth assessing the crack (time, cost, etc.)? Is there sufficient data to assess the crack? Will the crack grow? Is the growth mechanism understood and can it be quantified Will the assessment help to solve the underlying problem; i.e., will there be cracking elsewhere in the pipeline and will these cracks be growing?

form of damage. The cause of the cracking must be identified as part of the assessment, and indeed this is more important that the numerical side of the assessment. Cracking may be indicative of a problem with the design, the material, or the environment. In most cases, the purpose of the assessment is actually to prioritise the excavation of features. Then, the use of very complicated assessment methods may not be appropriate; screening methods may suffice. CONCLUSIONS 1. The pipeline-specific methods and generic methods both have a role in the assessment of crack-like defects in pipelines. 2. A comparison of predictions of full-scale tests on cracklike defects, and field and hydrotest failures shows that, for upper shelf steels, the pipeline-specific methods, such as the NG-18 equations, are more accurate than the generic methods, such as API 579-1 or BS 7910. The generic methods are conservative, but note they can give nonconservative predictions when used to calculate remnant fatigue lives. 3. The main cause of the conservatism in the generic methods is in their correlations between upper shelf Charpy V-notch impact energy and fracture toughness. 4. The pipeline-specific methods should be used when upper shelf behaviour is expected. The generic methods should be used when upper shelf behaviour cannot be proved. 5. The pipeline industry needs to develop better guidance for assessing crack-like defects in pipelines, which takes advantage of the benefits of the specific methods and the benefits of standardisation of the generic methods. 6. Cracks in a pipeline must be treated with caution. A crack is usually an indication that something is wrong with the pipelines material, environment, or operating conditions. An assessment will not provide an operator with this underlying cause of the cracking, and it is essential to understand the cause of the cracking before embarking on any assessment. REFERENCES
1. Anon., Guide to Methods for Assessing the Acceptability of Flaws in Metallic Structures, BS 7910 : 2005, British Standards Institution, London, UK, July 2005. Anon., Fitness-For-Service, API Recommended Practice 579-1 / ASME FFS-1 2007, Second Edition, The American Society of Mechanical Engineers, New York, USA, June 2007. Anon., Pacific Gas and Electric Company Natural Gas Transmission Pipeline Rupture and Fire San Bruno, California September 9, 2010, Pipeline Accident Report, NTSB/PAR11/01, PB2011-916501, Notation 8275C, National Transportation Safety Board, Washington, D.C., August 30, 2011. Woodley, D., 2011 The origin of intelligent pigs, Pipelines International, December 2011, p. 26-28. Eiber, R. J., Maxey, W. A., Bert, C. W., and McClure, G. M., 1981, The Effects of Dents on the Failure Characteristics of Linepipe, Battelle Columbus Laboratories, NG-18, Report No. 125, AGA Catalogue No. L51403.

The six steps in an assessment of crack-like defects in a pipeline are as follows: 1. 2. 3. 4. 5. 6. Understand the cause of the cracking. Quantify the size of the cracking, taking into account measurement accuracy and reliability. Quantify the material properties. Quantify the loads. Assess the crack. Solve the underlying problem, if there is one.

2.

3.

4. 5.

A Health Warning... Line pipe specifications and pipeline design codes do not permit cracks detected prior to commissioning to remain in a pipeline. They must be rectified (repaired) by dressing the crack or removing the defective section. Cracks are a severe

12

Copyright 2012 by ASME

6.

7.

8.

9.

10.

11.

12. 13.

14.

15.

16.

17.

18.

19.

20.

21.

22.

Anon., Manual for Determining the Remaining Strength of Corroded Pipelines, A Supplement to ASME B31 Code for Pressure Piping, ASME B31G-1984, The American Society of Mechanical Engineers, New York, USA, 1984. Kiefner, J. F., Duffy, A. R., Criteria for Determining the Remaining Strength of Corroded Areas of Gas Transmission Pipelines, Paper T, American Gas Association Operating Section on Transmission Conference, AGA, 1973, p. T86-T91. Shannon, R. W. E., Failure Behaviour of Line Pipe Defects, International Journal of Pressure Vessels and Piping, Vol. 2, 1974, p. 243-255. Jones, D. G., The Significance of Mechanical Damage in Pipelines, AGA-EPRG Linepipe Research Seminar IV, Duisburg, 1981. Anon., Guidance on some methods for the derivation of acceptance levels for defects in fusion welded joints, Published Document PD 6493 : 1980, British Standards Institution, London, UK, 1991. Anon., Guidance on methods for assessing the acceptability of flaws in fusion welded structures, Published Document PD 6493 : 1991, British Standards Institution, London, UK, 1991. Griffith, A. A., The Phenomena of Rupture and Flow in Solids, Philosophical Transactions, Series A, Vol. 221, 1920, p. 163-198. Irwin, G.R., Analysis of Stresses and Strains near the End of a Crack Traversing a Plate, Journal of Applied Mechanics, Vol. 24, 1957, p. 361-364. McClure, G. M., Eiber, R. J., Hahn, G. T., and Boulger, F. W., Research on the Properties of Line Pipe, American Gas Association, AGA Catalogue No. 40/PR, 1962. Duffy, A. R., Studies of Hydrostatic Test Levels and Defect Behaviour, Proceedings of the 3rd Symposium on Line Pipe Research, Pipeline Research Committee of the American Gas Association, Dallas, Texas, USA, AGA Catalogue No. L30000, 1965, p. 139-160. Rosenfield, A. R., Dai, P. K., and Hahn, G. T., Crack Extension and Propagation under Plane Stress, Paper A-14, Proceedings of the First International Conference on Fracture, T. Yokobori, T. Kawasaki, and J. L. Swedlow, Eds., Sendai, Japan, 12-17 September 1965, The Japanese Society for Strength and Fracture of Materials, Vol. 1, 1966, p. 223-258. Duffy, A. R., McClure, G. M., Maxey, W. A., and Atterbury, T. J., Study of the Feasibility of Basing Natural Gas Pipeline Operating Pressure on Hydrostatic Test Pressure, Report by Battelle Memorial Institute to the American Gas Association, AGA Catalogue No. L30050, 1968. Duffy, A. R., Hydrostatic Testing, Paper H, Proceedings of the 4th Symposium on Line Pipe Research, Pipeline Research Committee of the American Gas Association, Dallas, Texas, USA, AGA Catalogue No. L30075, 1969. Duffy, A. R., Eiber, R. J., and Maxey, W. A., Recent Work on Flaw Behaviour in Pressure Vessels, Paper M, Proceedings of Symposium on Fracture Toughness Concepts for Weldable Structural Steel, Risley, UK, 1969. Hahn, G. T., Sarrate, M., and Rosenfield, A. R., Criteria for Crack Extension in Cylindrical Pressure Vessels, International Journal of Fracture Mechanics, Vol. 5 No. 3, 1969, pp. 187-210. Kiefner, J. F., Fracture Initiation, Paper G, Proceedings of the 4th Symposium on Line Pipe Research, Pipeline Research Committee of the American Gas Association, Dallas, Texas, USA, AGA Catalogue No. L30075, 1969, p. G1-G36. Kiefner, J. F., Maxey, W. A., Eiber, R. J., and Duffy, A. R., Recent Research on Flaw Behavior During Hydrostatic Testing,

23.

24.

25.

26.

27.

28.

29.

30.

31.

32.

33.

34.

35.

American Gas Association, Operating Section Transmission Conference, 1971. Maxey, W. A., Kiefner, J. F., Eiber, R. J., and Duffy, A. R., Ductile Fracture Initiation, Propagation and Arrest in Cylindrical Vessels, ASTM STP 514, American Society for Testing and Materials, Philadelphia, 1972, p. 70-81. Kiefner, J. F., Maxey, W. A., Eiber, R. J., and Duffy, A. R., The Failure Stress Levels of Flaws in Pressurised Cylinders, ASTM STP 536, American Society for Testing and Materials, Philadelphia, USA, 1973, p. 461-481. Maxey, W. A., Fracture Initiation, Propagation and Arrest, Paper J, Proceedings of the 5th Symposium on Line Pipe Research, Pipeline Research Committee of the American Gas Association, Houston, Texas, USA, 1974. Kawaguchi, S., Hagiwara, N., Ohata, M., and Toyoda, T., Modified Equation to Predict Leak/Rupture Criteria for Axially Through-wall Notched X80 and X100 Line pipes Having Higher Charpy Energy, Paper No.: IPC04-0322, Proceedings of the 5th International Pipeline Conference, IPC 2004, Calgary, Alberta, Canada, 2004. Leis, B. N., and Zhu, X.-K., Leak vs. Rupture Boundary For Pipes With A Focus On Low Toughness and/or Ductility, Paper 21, 18th Biennial Joint Technical Meeting on Pipeline Research, APIA-EPRG-PRCI, San Francisco, California, USA, 16-19 May 2011. Leis, B. N., Brust, F. W., and Scott, P. M., Development and Validation of a Ductile Flaw Growth Analysis for Gas Transmission Line Pipe, Final Report to the Line Pipe Research Supervisory Committee of the Pipeline Research Committee of the American Gas Association, NG-18, Catalog No. L51543, June 1991. Leis, B. N., and Ghadiali, N. D., Pipe Axial Flaw Failure Criteria PAFFC, Version 1.0 Users Manual and Software, Topical Report to the Line Pipe Research Supervisory Committee of the Pipeline Research Committee of the American Gas Association, NG-18, Catalog No. L51720, May 1994. Jaske, C. E., and Beavers, J. A., Effect of Corrosion and StressCorrosion Cracking on Pipe Integrity and Remaining Life, MTI Publication No. 48, Proceedings of Second International Symposium on the Mechanical Integrity of Process Piping, Houston, Texas, USA, 30 January 1 February 1996, p. 287-297. Jaske, C. E., and Beavers, J. A., Review and Proposed Improvement of a Failure Model for SCC of Pipelines, Volume 1, Proceedings of Second International Pipeline Conference, IPC98, Calgary, Canada, American Society of Mechanical Engineers, 7-11 June 1998, p.439-445. Jaske, C. E., Development and Evaluation of Improved Model for Engineering Critical Assessment of Pipelines, Paper No.: IPC02-27027, Proceedings of IPC 2002: International Pipeline Conference, American Society of Mechanical Engineers, Calgary, October 2002. Dugdale, D. S., Yielding of Steel Sheets Containing Slits, Journal of the Mechanics and Physics of Solids, Vol. 8, 1960, p. 100-104. Bilby, B. A., Cottrell, A. H., and Swindon, K. H., The Spread of Plastic Yield from a Notch, Proceedings of the Royal Society, A272, 1963. Burdekin, F. M., and Stone, D. E. W., The Crack Opening Displacement Approach to Fracture Mechanics in Yielding Materials, Journal of Strain Analysis, Vol. 1, No. 2, 1966, p. 145153.

13

Copyright 2012 by ASME

36. Martin, M., Andrews, R. M., and Chauhan, V., Project #153L The Remaining Strength of Corroded Low Toughness Pipe, Final Report to the Department of Transportation Pipeline and Hazardous Materials Safety Administration, DTPH56-05-T0003, GL Noble Denton Report R9247 Issue 3.0, February 2009. 37. Garwood, S. J., Willoughby, A. A., and Rietjens, P., The Application of CTOD Methods for Safety Assessment in Ductile Pipeline Steels, Paper 22, International Conference on Fitness for Purpose Validation of Welded Constructions, The Welding Institute, London, UK, 17-19 November 1981. 38. Leis, B. N., Hydrostatic Testing of Transmission Pipelines: When It Is Beneficial and Alternatives When It Is Not, Prepared for the Design and Construction Committee of the PRCI, Report No. PR-3-9523, Catalogue No. L51844, December 2001. 39. Hosseini, A., Cronin, D., Plumtree, A., and Kania, R., Experimental Testing and Evaluation of Crack Defects in Line Pipe, Paper No.: IPC2010-31158, Proceedings of the 8th International Pipeline Conference, IPC 2010, Calgary, Alberta, Canada, September 27-October 1, 2010. 40. Rothwell, A. B., and Coote, R. I., A Critical Review of Assessment Methods for Axial Planar Surface Flaws in Pipe, International Conference on Pipeline Technology 2009, Ed. R. Denys, Ostend, Belgium, 12-14 October 2009. 41. Anon., Stress-Corrosion Cracking on Canadian Oil and Gas Pipelines, National Energy Board, Report of Inquiry, MH-2-95, 1996. 42. ASME B31G-2009 (Revision of ASME B31G-1991), Manual for Determining the Remaining Strength of Corroded Pipelines, A Supplement to ASME B31 Code for Pressure Piping, The American Society of Mechanical Engineers, New York, 2009. 43. Knauf, G., and Hopkins, P., The EPRG Guidelines on the Assessment of Defects in Transmission Pipeline Girth Welds, 3R International, 35, Jahrgang, Heft, 10-11, 1996, p. 620-624. 44. Denys, R. M., Andrews, R. M., Zarea, M., and Knauf, G., EPRG Tier 2 guidelines for the assessment of defects in transmission pipeline girth welds, Paper No.: IPC2010-31640, Proceedings of the 8th International Pipeline Conference, IPC 2010, American Society of Mechanical Engineers, Calgary, Alberta, Canada, September 27-October 1, 2010. 45. Cosham, A., Macdonald, K. A., Fracture Control in Pipelines Under High Plastic Strains : A Critique of DNV-RP-F108, IPC2008-64348, Proceedings of the 7th International Pipeline Conference, IPC 2008, American Society of Mechanical Engineers, Calgary, Alberta, Canada, September 30 - October 03, 2008. 46. Bates, N., Lee, D., and Maier, C., A Review of Crack Detection in-Line Inspection Case Studies, Paper No.: IPC2010-31114, Proceedings of the 8th International Pipeline Conference, IPC 2010, Calgary, Alberta, Canada, September 27-October 1, 2010.

14

Copyright 2012 by ASME

You might also like